The Optimization of Rifle Barrel Harmonics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 72

University of Vermont

UVM ScholarWorks

Graduate College Dissertations and Theses Dissertations and Theses

2022

The Optimization of Rifle Barrel Harmonics


Daniel O'Neil
University of Vermont

Follow this and additional works at: https://scholarworks.uvm.edu/graddis

Part of the Mechanical Engineering Commons

Recommended Citation
O'Neil, Daniel, "The Optimization of Rifle Barrel Harmonics" (2022). Graduate College Dissertations and
Theses. 1608.
https://scholarworks.uvm.edu/graddis/1608

This Thesis is brought to you for free and open access by the Dissertations and Theses at UVM ScholarWorks. It
has been accepted for inclusion in Graduate College Dissertations and Theses by an authorized administrator of
UVM ScholarWorks. For more information, please contact schwrks@uvm.edu.
The Optimization of Rifle Barrel Harmonics

A Thesis Presented

by

Daniel O’Neil

to

The Faculty of the Graduate College

of

The University of Vermont

In Partial Fullfillment of the Requirements


for the Degree of Master of Science
Specializing in Mechanical Engineering

August, 2022

Defense Date: July 19, 2022


Thesis Examination Committee:

Dryver Huston, Ph.D., Advisor


Eric Hernandez, Ph.D., Chairperson
Jihong Ma, Ph.D.
Cynthia J. Forehand, Ph.D., Dean of the Graduate College
Abstract

From the optimization of mechanical systems perspective, firearms offer a plethora of complex
scenarios. Many variables, coupled with sometimes unstable systems, must be balanced appro-
priately by the user to achieve precision, that is to say, to hit the target bullseye from some
distance away with a small projectile launched on a parabolic trajectory. The sources of error
entirely affect final terminal precision, some of them related to human factors, others related to
the quality of the ammunition, and the remaining associated with the mechanical capability of
the weapon. This thesis documents the vibrations analysis of an isolated rifle barrel, identifies
barrel tip displacement and angle of rotation in firing scenarios both analytically and empirically,
and makes recommendations on optimizations and control schemes to improve overall precision
capability.
Table of Contents

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

2 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

3 Barrel Loads and Natural Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . 11

4 Analytical Modeling Results Using ANSYS . . . . . . . . . . . . . . . . . . . . 28

5 Empirical Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

6 Barrel Optimizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

6.1 Barrel Length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

6.2 Barrel Moment of Inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

6.3 Modal Mass Tuning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

8 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

8.1 Barrel Vibrations MATLAB Code . . . . . . . . . . . . . . . . . . . . . . . . . . 55

8.2 MATLAB Code Output Example . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

ii
List of Figures

1 Tikka T3x Bolt Action Rifle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

2 Cartridge Loaded in Rifle Chamber . . . . . . . . . . . . . . . . . . . . . . . . . . 3

3 Typical Barrel Rifling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

4 Representative Pressure Curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

5 A. Mallock Rifle Modes from ”Vibration of Rifle Barrels”, 1901 . . . . . . . . . . 6

6 Carbon Fiber Wrapped Machine Gun Barrels, Pyka et al., 2019 . . . . . . . . . . 7

7 Tank Armored Fighting Vehicle Dynamic MRS, Dursun et al., 2017 . . . . . . . 8

8 G. Kolbe Tuned Rifle to Optimize Projectile Exit Time, 2021 . . . . . . . . . . . 9

9 TacomHQ Structured Rifle Barrel . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

10 Chamber and Bore Forces During Firing Event . . . . . . . . . . . . . . . . . . . 11

11 20 DOF Beam With Displacement Reactions at Node 1 . . . . . . . . . . . . . . 13

12 20 DOF Beam With Displacement Reactions at Node 2 . . . . . . . . . . . . . . 13

13 Displacement Degrees of Freedom Stiffness Matrix Through Node 10 . . . . . . . 14

14 Rotational Degree of Freedom Node 11 . . . . . . . . . . . . . . . . . . . . . . . . 14

15 Rotational Degree of Freedom Node 12 . . . . . . . . . . . . . . . . . . . . . . . . 15

16 20 Degree of Freedom Stiffness Matrix K, for Tikka T3X Rifle . . . . . . . . . . . 16

17 Condensed Stiffness Matrix, [Kt ] . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

18 Discretized Mass Matrix, [M ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

19 Natural Frequencies of Tikka T3X System, Eigenvalues . . . . . . . . . . . . . . . 19

20 Modeshapes of Tikka T3X System, Eigenvectors . . . . . . . . . . . . . . . . . . 20

21 Modal Mass Matrix for 20DOF Tikka System . . . . . . . . . . . . . . . . . . . . 23

22 Impulse Response for Tikka System . . . . . . . . . . . . . . . . . . . . . . . . . 24

23 Tikka System Response, Mode 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

24 Tikka System Response, Mode 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

25 ANSYS Static Structural Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

26 ANSYS Static Structural Moving Force, Step 1 . . . . . . . . . . . . . . . . . . . 29

iii
27 ANSYS Modal Analysis of Tikka T3X System, Mode 1, 38.6 Hz . . . . . . . . . . 29

28 ANSYS Modal Analysis of Tikka T3X System, Mode 2, 240.7 Hz . . . . . . . . . 30

29 ANSYS Modal Analysis of Tikka T3X System, Mode 3, 669.8 Hz . . . . . . . . . 30

30 Tikka System Rifle Barrel Test, DEWESoft DAQ, PCB Accels . . . . . . . . . . 32

31 Butterworth Filter Response Depending on Filter Order . . . . . . . . . . . . . . 33

32 Tikka System Rifle Barrel Test, Tap Test, 100kHz . . . . . . . . . . . . . . . . . 33

33 Tikka System Rifle Barrel Test, Accelerometer Placement . . . . . . . . . . . . . 34

34 Tikka System Rifle Barrel Test, Time Series Acceleration Data, Test 1, Shot 1 . . 35

35 Tikka System Rifle Barrel Test, Empirically Determined Modes, Test 1, Shot 1 . 36

36 Tikka System Rifle Barrel Test, Time Series Acceleration Data, Test 1, Shot 2 . . 36

37 Tikka System Rifle Barrel Test, Empirically Determined Modes, Test 1, Shot 2 . 37

38 Saturated Accelerometer, Test 2, Shot 2 . . . . . . . . . . . . . . . . . . . . . . . 38

39 Tikka System Rifle Barrel Test, Time Series Acceleration Data, Test 2, Shot 1 . . 39

40 Tikka System Rifle Barrel Test, Time Series Acceleration Data, Test 2, Shot 2 . . 40

41 Short Barrel Optimization vs. Tikka T3X System Response . . . . . . . . . . . . 43

42 Short Barrel Optimization vs. Tikka T3X System Action Time Response . . . . 44

43 12.5” Mid Gas System in 5.56mm (L) and 12.5” Piston System in 5.56mm (R) . 45

44 Grid Style Extruded Barrel Blank Model, Area Moment of Inertia Optimization 46

45 Grid Style Extruded Barrel Blank, Modal Analysis, Mode 2 . . . . . . . . . . . . 47

46 Moment of Inertia Optimization vs. Tikka T3X System Response . . . . . . . . . 48

47 Moment of Inertia Optimization vs. Tikka T3X System Action Time Response . 49

48 Modal Mass Tuning Scenario vs Nominal Tikka T3X Vibration Response . . . . 51

49 Grid Style Extruded Barrel Blank, Modal Analysis, Mode 2 . . . . . . . . . . . . 52

iv
List of Tables

1 Displacements from Empirical Test Results (* denotes data in question) . . . . . 40

2 Tikka T3X Baseline Muzzle Performance at Projectile Exit . . . . . . . . . . . . 41

3 Tikka T3X System L/2 Barrel Performance . . . . . . . . . . . . . . . . . . . . . 42

4 Optimization of Grid Style Extruded Barrel . . . . . . . . . . . . . . . . . . . . . 46

5 Tikka T3X System Tuned Nodal Masses Performance . . . . . . . . . . . . . . . 51

v
1 INTRODUCTION 1

1 Introduction

The use of tools to more efficiently perform a task, with few exceptions, is a decidedly human trait.

Throughout history, humans have continued to invent, improve, design, iterate, and otherwise

engineer tools to achieve goals and broaden overall technological capability. Of particular interest

is in optimizing the modern rifle barrel for the purposes of improving precision and thereby

augmenting overall mechanical performance of the weapon and ammunition combination. As

a reader who is skilled in the art may understand, a mechanical assembly such as a rifle is

susceptible to the laws of dynamics just like any other mechanism, and there are inefficiencies

and errors that sum up to an overall degradation in performance. It is in eliminating, redirecting,

or minimizing these sources of error that it is possible to achieve greater precision in a mechanical

system such as a rifle. This thesis explores the mathematical options for optimizing rifle barrel

performance through minimizing error. One source of error is the way the barrel behaves during

a firing event. Consider the rifle(1) shown in Figure 1 in caliber 6.5 Creedmoor, which is a

6.5mm diameter projectile traveling at 2650 feet per second and with a weight of 130 grains or

9.07 grams.
1 INTRODUCTION 2

Figure 1: Tikka T3x Bolt Action Rifle

This is an example of a modern bolt action rifle, and is representative of most commercially

available sporting rifles. The rifle barrel itself is ”free floated”, meaning that it does not contact

the firearm stock at any point along its length from the action to the muzzle. This degree of

separation ensures the barrel can vibrate freely and naturally, without influence by external forces

through stock contact, and perform repeatably for each firing event.

When a cartridge is loaded into the rifle and fired through the barrel as designed, the firing pin

hits the cartridge primer, which ignites the smokeless powder (propellant) internally contained in

the cartridge. The rapid burning of the propellant increases pressure severely until the projectile

is forced out of the cartridge and down the barrel at a predetermined velocity based on projectile

mass, propellant load, and cartridge volume (Figure 2).


1 INTRODUCTION 3

Figure 2: Cartridge Loaded in Rifle Chamber

(2)

As the projectile is forced out of the cartridge and into the rifling (Figure 3), the rifling

engraves the projectile body and imparts spin. Barrels are typically designed with a consistent

twist rate e.g. 1 rotation in 8 inches of travel or 1:8. With larger cannon barrels, it is common

to have a variable twist rate or ”gain twist”, where the projectile spin rate increases the further

it travels down the barrel and reaching a max spin at the muzzle end. This paper will consider

only barrels with consistent twist rates, although the same principle could be applied to gain

twist barrels.

Figure 3: Typical Barrel Rifling

(3)
1 INTRODUCTION 4

Optimal barrel twist rate for a given projectile may be calculated using the Green Hill For-

mula, given bullet diameter, length, specific gravity, and constant C (C=150 for projectile veloc-

ities up to 2,800 feet per second(3)).

r
C ∗ D2 Sg
T = ∗ (1)
L 10.9

Utilizing data for the 6.5 Creedmoor, Equation 1 becomes(3):

r
150 ∗ .2642 10.05 inches
T = ∗ = 7.27 (2)
1.380 10.9 rotation

For this projectile, a 1:7.27 twist rate would be optimal, however, commercially, barrels are

produced in a 1:8 twist rate to handle a larger variety of available projectiles weights. While

the projectile is traveling down the barrel, it is influenced by the burn rate of the propellant,

and the subsequent pressure developed during the firing event. This it typically captured in

a pressure curve, which peaks while the projectile is still traveling down the barrel bore. A

representative pressure cure is shown below in Figure 4(4). It should be noted that the projectile

is only influenced by barrel movement while it is inside the rifle barrel. After it exits, the barrel

may move and vibrate extensively however, it is irrelevant from the perspective of the projectile

and its intended target.


2 LITERATURE REVIEW 5

Figure 4: Representative Pressure Curve

For the short time that the projectile travels within the barrel bore, it is subjecting the barrel

to reactionary loads, and follows the movement of the barrel as it vibrates at its natural modes.

Loads reacted by the barrel, especially for imperfect projectiles or barrels, and natural vibration

will cause deviations between the point of aim, and the point of projectile impact.

2 Literature Review

An extensive amount of work has already been accomplished on rifled barrels for both commercial

and military purposes. Mr. J. Siewert points out that ”Barrel-pointing variability accounts for

85%-95% of the barrel-related dispersion error budget...”(21). As far back as the early 1900s,

engineers and technicians were working to understand the dynamic movement of a rifle barrel

and indeed, of the weapon itself, for the purposes of improving precision and system design.

On June 6, 1901, Mr. A. Mallock by way of Lord Rayleigh, F.R.S., communicated a paper

that discussed the phenomenon of rifle ”flip” attributed to the method of which a rifle barrel is

attached to its stock. That is to say, the rifle barrel center axis vertical offset from the point
2 LITERATURE REVIEW 6

at which the stock contacts the user’s shoulder, causing the muzzle to offset from the aimpoint

during a firing (recoil) event. Mr. Mallock proposed rifle mode shapes based on the firing

excitation, derived rudimentary equations of motion and that which described response, and

measured natural frequencies based on a Lee Enfield Rifle(12). Mallock concludes that muzzle

tip slope at the moment of projectile exit and cartridge velocity variation as having a distinct

effect on weapon precision. Mallock expresses in his paper that a ”good modern sporting rifle

will shoot within a probable deviation of considerably less than 2 feet from the intended path.”

Modern weapons and ammunition have made great strides since then!

Figure 5: A. Mallock Rifle Modes from ”Vibration of Rifle Barrels”, 1901

More modern studies tend to be of the form of a Finite Element Analysis (FEA), checked

against classical undamped equations of motion. Xu, Guan, Liu, and Xu in their paper ”Study

on Barrel Vibration Characteristics of Typical Sniper Rifle”, use a traditional FEA approach to
2 LITERATURE REVIEW 7

analyze a sniper rifle barrel, develop modeshapes and displacement, and then compare results to

empirically collected data using a high speed camera. Their model to empirical data correlation

was roughly 80% (13).

This was similarly accomplished by Pyka, Bocian, Jamroziak, Kosobudzki, and Kulisiewicz in

the paper ”Concept of a Gun Barrel Based on the Layer Composite Reinforced with Continuous

Filament”. In this case, the team developed an FEA model of a subject machine gun barrel,

collected test data on that barrel, and then proposed a carbon fiber exterior sheath with which

to augment barrel stiffness and therefore shift natural frequency in a mechanically favorable

direction (14).

Figure 6: Carbon Fiber Wrapped Machine Gun Barrels, Pyka et al., 2019

˙ ELEK having to do with the dynamics


Several works have come from F irat B Ü Y Ü KC IV
˙ ELEK’s
and control of large bore cannons, specifically used in armored vehicles. In B Ü Y Ü KC IV

work on ”Analysis and Control of Gun Barrel Vibrations, he delves into the factors affecting

tank barrel motion and types of vibrations sustained in an armored vehicle. His data includes

accelerometer data collected from live tank operations and from a fixtured shaker armature to

support his hypothesis. Model development consistently utilizes FEA to approximate a real

world system, and develop mode shapes and displacements. His solution is a ”Tuned Vibra-

tion Absorber”, or oscillating single degree of freedom system that cancels or reduces vibration
2 LITERATURE REVIEW 8

amplitudes at individual barrel modes (15).

He also published a second work titled ”A Review on the Gun Barrel Vibrations and Control

for a Main Battle Tank” in cooperation with Dursan and Utlu, where they explored the relation-

ship of tank cannon muzzle length to its anticipated displacement. He finds a direct relationship

between barrel length and accuracy of that combat vehicle, especially for vehicles that are not

static during a firing event. Their solution was a Dynamic Muzzle Reference System (DMRS)

that controlled aiming offset and stabilized muzzle vibrations during the event to maximize pre-

cision of the cannon barrel. They also recommend shorter gun barrels (relative stiffness) and

active vibration controls systems as potential optimization opportunities (16).

Figure 7: Tank Armored Fighting Vehicle Dynamic MRS, Dursun et al., 2017

Similarly, Koc, Esen, and Cay from Sakarya University published a paper on large caliber

cannon barrels utilized in armored vehicles, specifically the effect of unbalanced projectiles and

their related error. In their paper, they again define equations of motion particular to a tank

cannon for each axis of oscillation, generate modal mass, stiffness matrix, and damping relative

to barrel elements, and then generate cannon barrel predicted response. The purpose of their

effort seems to be the utilization and validation of a novel analysis tool named 12 DOF FEM

and Newman â algorithm (17).

Mr. G. Kolbe developed an analysis and method for tuning rifle barrels to an optimum

point for precision shooting at a known distance in his paper ”Using Barrel Vibrations to Tune
2 LITERATURE REVIEW 9

a Barrel”. After a review of prior work both reputed and infamous (to include referencing

Mr. Mallock’s work from 1901 above), he discusses measurement of barrel motion in one axis

(vertical), and the measurement of the exact moment the projectile leaves the muzzle by using

what he terms a ”muzzle gate”. In this way, Kolbe collects muzzle angle, and muzzle velocity,

which previously had only been derived using position, velocity, and acceleration mathematical

models. He concludes with demonstrations of firing events with an untuned barrel firing groups

at 50 yards, and a barrel tuned with a mass at the end to control projectile exit timing. With

consistent projectile exit, Kolbe is able to significantly increase rifle precision. It is apparent that

this tuning must be accomplished any time ammunition load or projectile weight change (18).

Figure 8: G. Kolbe Tuned Rifle to Optimize Projectile Exit Time, 2021

The company TacomHQ has developed a product termed ”Structured Barrels” that they

market as harmonically dead (19). The actual product is a large diameter barrel blank that is

gun drilled and rifled along the center axis, as a normal rifle barrel would be, but it is then gun

drilled concentrically to form tubes around the barrel. TacomHQ offers some explanation and

analysis to support their claim. It appears that they have optimized a geometric shape, as it

relates to its cross sectional moment of inertia, for increased stiffness, lower weight, and a natural

frequency shift to reduce muzzle degree of freedom displacement and rotation.


2 LITERATURE REVIEW 10

Figure 9: TacomHQ Structured Rifle Barrel

There is also a product sold under the company ”Recreational Software, Inc”, that allows a

user to measure exact moment of exit of the projectile from a rifle barrel. This software guides the

user in developing customized reloaded cartridges that take advantage of a specific rifle’s barrel

behavior during a firing impulse. It refers to muzzle ”whip” and barrel ”ringing” to describe

barrel vibrations as a result of the firing event (22).

In Mr. Potts’ ”Things that affect rifle accuracy (and how to correct them)” article for Shootin-

gUK, he discusses rifle vibrations as a significant source of error. ”The cause of more inaccuracies

and barrel problems than realized. The barrel vibrates as the bullet passes along its length, so

it is best to keep that vibration as smooth and uniform as possible.” He advocates for different

types of rifle action bedding, that is, a way to stabilize and support the rifle action with sub-

stances like epoxy. He also mentions the importance of having a free floated barrel that is not

contacted anywhere by the rifle fore end (23).


3 BARREL LOADS AND NATURAL VIBRATIONS 11

3 Barrel Loads and Natural Vibrations

Referencing Calucci and Jacobsen(5), a perfectly homogenous and symmetric barrel will expe-

rience a predictable array of loads as the projectile travels down the bore. ”During and after

firing, the unbalanced forces on the gun may be categorized as the gas force FR , the projectile

resistance force FP r , and the rifling force FT ”(5). Figure 10 below shows where these loads occur.

Figure 10: Chamber and Bore Forces During Firing Event

The rifling force may be related to the projectile force via the following equation:

k
FT = ( d )2 ∗ FP r ∗ tan(α) (3)
2

Where k is the projectile radius of gyration, d is projectile diameter, and α is the angle of the

rifling. Notionally, for a perfect rifle barrel firing a perfect projectile, there is no force unbalance

reacted by the barrel, and thus no real impact to overall system precision at the moment the

projectile exits the muzzle of the barrel and thereby out of its influence. With an asymmetric

rifled bore, it is apparent that a force component of some magnitude would act on the rifle barrel

in such a way as to effectively generate a moving load traveling at the point of projectile rifling

contact perpendicular to the barrel wall. This force would be the rifling force FT at its peak due

to the barrel imperfection.

Let us shelve this concept for a moment and explore the mathematics that govern barrel

movement during a firing event. For the purposes of developing a model, the barrel is assumed
3 BARREL LOADS AND NATURAL VIBRATIONS 12

to behave elastically such that the equations of motion for a beam may be used. In this case, the

boundary conditions are for a fixed-free configuration, and will be stated plainly below.

Firstly, let us start with the general equation of motion for a classically damped system:

mẍ + cẋ + kx = F (t) (4)

Equation 4 above is for a single degree of freedom system, subjected to a forcing function on

the right side of the equation. Although only for a single degree of freedom, this equation may

be leveraged to analyze n degrees of freedom (DOFs) through the process of Modal Analysis.

In this case a more appropriate rendition of the equation for n degrees of freedom is the matrix

format:

[M ] {ẍ} + [C] {ẋ} + [K] {x} = {F (t)} (5)

Where the [M ] is the DOF lumped mass matrix, [C] is the damping matrix, and [K] is the

stiffness matrix for the system. System free response and eigenvalues are determined from the

mass and stiffness matrix only.

Considering our example system, the Tikka T3X 6.5 Creedmoor rifle, we can develop a math-

ematical model for barrel position relative to an axis coincident and parallel with the bore

centerline by applying the classical equations of motion to a discretized number of segments of

barrel length. In this way, each nodal displacment motion may be quantified.

It is important to select an adequate number of degrees of freedom for this method of calcu-

lation, as too few will result in noticeable error from the discretization. The first iteration for

this paper only contained 6 DOF and exhibited approximately 15-20% error when compared to

Finite Element Analysis results. Let us consider 10 nodes and thus 20 degrees of freedom, un-

derstanding that each node will exhibit displacement and rotation. Utilizing beam displacement

and rotation shapes from Tedesco (6), it is possible to develop the stiffness matrix one degree of

freedom at a time. Degree of Freedom 1 and reaction moments and loads are displayed below in
3 BARREL LOADS AND NATURAL VIBRATIONS 13

Figure 11.

Figure 11: 20 DOF Beam With Displacement Reactions at Node 1

This is possible by applying a displacement of magnitude ”1” at the node of interest, and

fixing displacement of all other nodes in the system to 0. At this point, simple beam equations

apply at node 1. The end result is the stiffness vector {K1 }, which is a column vector of length

20 (for each DOF), and is displayed below:

 
24EI −12EI 6EI
K1 = , , 0, 0, 0, 0, 0, 0, 0, 0, 0, 2 , 0, 0, 0, 0, 0, 0, 0, 0 (6)
L3 L3 L

Likewise for degree of freedom 2 at Node 2, the corresponding beam shape and stiffness vector

{K2 } is:

Figure 12: 20 DOF Beam With Displacement Reactions at Node 2

 
−12EI 24EI −12EI −6EI 6EI
K2 = , , , 0, 0, 0, 0, 0, 0, 0, , 0, 2 , 0, 0, 0, 0, 0, 0, 0 (7)
L3 L3 L3 L2 L

This pattern continues through degree of freedom 10, which completes the displacement
3 BARREL LOADS AND NATURAL VIBRATIONS 14

lbs
shapes of each node. For the Tikka T3X barrel, utilizing Esteel = 30 ∗ 106 in 2 , OD = .75in,

π 4
ID = .22in, I = 64 (OD − ID4 )in4 and L = 23.5/n inches, the first 10 DOFs for the stiffness

matrix are shown below. Note that the first 10 DOFs are displacement only, and the final stiffness

matrix will be a 20 x 20.

Figure 13: Displacement Degrees of Freedom Stiffness Matrix Through Node 10

It should also be noted that the stiffness matrix is always symmetric. After applying beam

formulas for each of the displacement degrees of freedom, the same must then be repeated for

rotations at each node for 11-20th degrees of freedom. Similar to the displacements applied above

to generate stiffness values, now rotations are applied at each node and reaction moments and

forces are calculated. All other nodes have rotation fixed at 0, and the node of interest has a

rotation (moment) applied. For node 11, the subsequent beam rotation shape(6) and reactions

are:

Figure 14: Rotational Degree of Freedom Node 11


3 BARREL LOADS AND NATURAL VIBRATIONS 15

And the resulting stiffness vector {K11 } is:

 
−6EI 8EI 2EI
K11 = 0, , 0, 0, 0, 0, 0, 0, 0, 0, , , 0, 0, 0, 0, 0, 0, 0, 0 (8)
L2 L L

This process continues on to node 12 in the same fashion, generating the beam rotation shape

as follows:

Figure 15: Rotational Degree of Freedom Node 12

and resulting stiffness vector {K12 }:

 
6EI 6EI 2EI 8EI 2EI
K12 = , 0, , 0, 0, 0, 0, 0, 0, 0, , , , 0, 0, 0, 0, 0, 0, 0 (9)
L2 L2 L L L

Progressing through each displacement and rotational degree of freedom allows the develop-

ment of the system stiffness matrix, dependent only on E, I, and beam length L. Beam length

is the linear distance between nodes on the beam, and is the resolution of the discretization that

we are performing on the beam.

The resulting overall stiffness matrix with all degrees of freedom for 10 nodes is:
3 BARREL LOADS AND NATURAL VIBRATIONS 16

Figure 16: 20 Degree of Freedom Stiffness Matrix K, for Tikka T3X Rifle

Going back to the classical equation of motion below, we can now plug in the stiffness matrix,

leaving the damping matrix [C] and the mass matrix [M ] still undefined. The mass matrix is

simple in that it is a diagonal matrix with each mii value equal to the overall mass of the barrel

divided over the number of discretized segments. The damping matrix [C] is normally estimated

during mathematical modeling, or determined empirically through test, which will be covered

later in this thesis. For the purposes of this mathematical model, the damping value was assumed

to be 5% for each degree of freedom.

[M ] {ẍ} + [C] {ẋ} + [K] {x} = {F (t)} (10)

Because rotational degrees of freedom do not have a mass associated with them, they may be

disregarded by conducting a Static Condensation of the stiffness matrix K. It follows that the

stiffness matrix [K] may be dissected into its displacement and rotational components per(6):

      
 Ktt Ktθ   X   Ft   Ft 
  = = 
Kθt Kθθ θ Fθ 0

Where K is defined as:

 
 Ktt Ktθ 
 =K
Kθt Kθθ
3 BARREL LOADS AND NATURAL VIBRATIONS 17

The subvectors X and θ correspond to translational and rotational displacement(6) at each

degree of freedom. The force subvectors Ft and Fθ correspond to the displacement subvectors X

and θ. If force subvectors in the system do not include rotational components, then Fθ = 0, and

the stiffness matrix may be condensed. In this case, evaluating the rotational displacements per

the above, a second submatrix may be developed as(6):

−1
θ = −Kθθ Kθt X (11)

Expanding the submatrix results in:

Ktt X + Ktθ θ = Ft (12)

And substituting Equation 11 into Equation 12 results in:

−1
(Ktt − Ktθ Kθθ Kθt )X = Ft (13)

Therefore, the overall stiffness matrix equation shown above may be represented more simply

as the condensed matrix:

Kt X = Ft (14)

Where Kt is represented as:

−1
Kt = Ktt − Ktθ Kθθ Kθt (15)

In this way, only the degrees of freedom with associated mass are represented in the condensed

matrix (10x10) which also correlates to the sizes of both [M ] and [K]. This extensive matrix

math is easily computed in MATLAB, resulting in the following:


3 BARREL LOADS AND NATURAL VIBRATIONS 18

Figure 17: Condensed Stiffness Matrix, [Kt ]

Figure 18: Discretized Mass Matrix, [M ]

With [K] and [M ], we are now interested in the modeshapes and natural frequencies of the

system. This is achieved by solving the Eigenproblem for both Eigenvectors and Eigenvalues.

Knowing the equations of motion for the system (Equation 4), and disregarding any system

damping, we can set it equal to zero and postmultiply by the inverse of the mass matrix ([M ]−1 )

to yield (6):

[I] {ẍ} + [Kt ][M ]−1 {x} = 0 (16)

Since:
3 BARREL LOADS AND NATURAL VIBRATIONS 19

[D] = [M ]−1 [Kt ] (17)

And we know that simple harmonic motion takes the form:

{x(t)} = {A} sin(ωt + φ) (18)

then:

[[D] − λ[I]] {A} = 0 (19)

Which is recognized as the standard eigenvalue problem. Its solution results in an eigenvalue

diagonal matrix of natural frequencies for each degree of freedom, and an eigenvector matrix of

modeshapes. The solution to which is simplified by utilizing the eig function in MATLAB, and

resulting in the following frequency and modeshape matrices:

Figure 19: Natural Frequencies of Tikka T3X System, Eigenvalues


3 BARREL LOADS AND NATURAL VIBRATIONS 20

Figure 20: Modeshapes of Tikka T3X System, Eigenvectors

Although the stiffness matrix and corresponding Eigenvalue problem was developed indepen-

dent of the forcing function F (t), we now must consider how damping and the forcing function

impact the system’s vibration response during a firing event. If the forcing function remained

static at one point of the barrel, the solution and subsequent system response would be trivial.

However, the projectile in this case is not stationary. The forcing function, for the purpose of

this mathematical model, was set up as an asymmetric load on the rifling, traveling down the

barrel with the projectile. In order to correctly build this model, Warburton was consulted on

moving loads in vibrational systems(7), and Blevins was used for modeshape equations(8).

At t = 0, it is assumed that the projectile is still unfired and in the chamber, and no motion

has yet started. At t corresponding to the travel of the projectile down the barrel distance x at

velocity U , the subsequent projectile position is x = U t. If said projectile travels at a muzzle

velocity of 2, 650 fsec


eet
, or 31, 800 inch
sec , then the projectile is influenced by barrel vibrations for the
L
first U of the response. For the Tikka, with a barrel length of 23.5 inches, the time of influence

is:

L 23.5in
ti = = in
= .00074sec (20)
U 31800 sec

However, this erroneously assumes an instantaneous velocity, and does not account for the

realistic time for primer strike, combustion and pressure to build, and projectile exit from the
3 BARREL LOADS AND NATURAL VIBRATIONS 21

muzzle. This term is called action time, and will be assumed to be 0.002 seconds, or a corre-
in
sponding average velocity of 11, 750 sec , in line with similar cartridge industry empirical data.

Warburton sets up the moving forcing function (constant force) for a simply supported beam

as(7):

1 sπU T
φs (a) = 2 2 ∗ sin (21)
l

Where s is the mode (s=1, 2, 3,...), U is projectile velocity, and t is time of projectile travel
1
(starting from 0). The 2 2 is a normalizing factor particular to his solution for a moving force on

a simply supported beam(7). Warburton provides the solution as:

1 P sin(sπ Ul t )
qs = Bs sin(ωs t) + Cs cos(ωs t) + 2 2 (22)
m ωs2 − (sπ Ul t )2

If the beam starts at rest when t=0, then initial conditions and boundary conditions drive

q(0) = 0 = q̇ and the solution becomes (7):

1 Ut U
2 2 P sin(sπ l ) − (sπ lωs )sin(ωs t)
qs = (23)
m ωs2 − (sπ Ul )2

Since gunfire provides broadband excitation, we assume that the forcing function occurs at

the natural frequency of the mode being excited in the Tikka T3X system. Equation 23 may

then be replaced by:

1
22 P
q¨s + ωs2 qs = sin(ωs t) (24)
m

With the conditions at t=0 applied, the solution becomes:

P
qs = 1 (sin(ωs t) − ωs tcos(ωs t)) (25)
2 mωs2
2

However, this is inadequate to describe the function for a cantilever beam such as a rifle

barrel; instead the case needed is fixed-free. Referring to Blevins modeshapes, it is understood
3 BARREL LOADS AND NATURAL VIBRATIONS 22

that a fixed-free system’s (Blevins labels it a ”clamped-free” system) modeshapes may be math-

ematically re-created using the following sine and cosine functions:

λi x λi x λi x λi x
ρi = cosh( ) + cos( ) − σi [sinh( ) + sin( )] (26)
L L L L

Where λi specifically for Equation 26, is the following vector, corresponding to mode(8):

λi = {1.87510407, 4.69409113, 7.85475744, ...} (27)

And likewise, σi is a vector corresponding to each mode(8):

σi = {.734095514, 1.018467319, .999224497, ...} (28)

Because we know that Warburton’s method in Equation 24 is based on a simply supported

beam, we also know the modeshapes are sinusoidal. For a fixed-free system, the sin in Equation

24 was replaced with the modeshape equation in Equation 26, and then utilized to derive the

modal force equation as:

1
2 2 P ρi
M Fi = (29)
µii

In the modal force equation above (Equation 29), P is the magnitude of the moving force, ρi is

the modal influence (Blevins modeshape), and µii is the modal mass value for the ith node. The

modal mass matrix [µ] is necessary to achieve proportional values in the response calculation,

and is a result of the transformation of the form:

T
[µ] = {ψ} [M ] {ψ} (30)

The actual Modal Mass matrix used is:


3 BARREL LOADS AND NATURAL VIBRATIONS 23

Figure 21: Modal Mass Matrix for 20DOF Tikka System

At this point, the moving force for each mode, M Fi , may be calculated, which results in the

right side of the modal equation shown above. Set up for the Tikka T3X fixed-free system, the

response equation is:

1
2 2 P ρi
q¨s + ωs2 qs = = M Fi (31)
µii

This modal force is applicable only while the projectile is traveling down the barrel. At all

other points in time, the modal force on the rigth side of the equation is 0. We know that this

modal equation of motion was derived using a modal transformation of the form:

[Ψ]T [M ][Ψ] {q̈} + [Ψ]T [C][Ψ] {q̇} + [Ψ]T [K][Ψ] {q} = M Fi (32)

The next question is how to solve Equation 31? A direct integration is possible, but cumber-

some and nontrivial. It is more appropriate to solve using the convolution integral or Duhamel’s

integral. Duhamel’s integral is stated as(6):

dx = F (τ )dτ ∗ h(t − τ ) (33)

Due to the principle of superposition, meaning that each segment of the the response inte-

gration may be added together to obtain the complete response, the Duhamel integral may be

expressed instead as:


3 BARREL LOADS AND NATURAL VIBRATIONS 24

Z t
x(t) = F (τ )h(t − τ )dτ (34)
0

In order to solve Equation 31, we must utilize Duhamel’s integral to convolve the impulse

response of the system h(t) with the modeshape matrix ψ.

The impulse response h(t) is easily calculated from the characteristics of the dynamic system

ω or natural frequency, ωd or damped natural frequency, mass, and time. It is:

1 −ζωt
h(t) = e sin(ωd t) (35)
mωd

Figure 22: Impulse Response for Tikka System

Because the principle of superposition was used to establish Duhamel’s integral, we may apply
3 BARREL LOADS AND NATURAL VIBRATIONS 25

it to this linear system and determine the response due to excitation from a gunfire event(6).

Substituting the impulse response into Duhamel’s integral gives:

Z t
1
x(t) = F (τ )e−ζω(t−τ ) sinωd (t − τ )dτ (36)
mωd 0

The result of evaluating this convolution integral, and multiplying by the modal force, is in

modal coordinates and must be transferred back to physical coordinates by multiplying by the

modeshape matrix ψ which gives the equation of motion q for each mode. Again due to the

principle of superposition, the total response may be obtained by adding up the q value for each

mode at each point in time.

This process was accomplished for modes one and two in this mathematical model, and the

resulting responses are shown below:


3 BARREL LOADS AND NATURAL VIBRATIONS 26

Figure 23: Tikka System Response, Mode 1


3 BARREL LOADS AND NATURAL VIBRATIONS 27

Figure 24: Tikka System Response, Mode 2

The first three modes of the Tikka system occur at 35.7, 224.8, and 632.2 Hz respectively

according to the discretized model discussed above.

A double check utilizing the Blevins calculations for fixed-free beam natural frequency result

in:

λi EI 1
fi = ( )2 (37)
2πL m

Where λi = [1.87510407, 4.69409113, 7.85475744, ...] according to the values specified in Blevins

Modeshapes(8).

The Blevins calculation results in the first three modes occurring at 39.3, 246.0, and 688.9

Hz which are within 10% of the discretized model


4 ANALYTICAL MODELING RESULTS USING ANSYS 28

4 Analytical Modeling Results Using ANSYS

ANSYS was also utilized to correlate the analytical models to the subsequent empirical data. For

the ANSYS model, the rifle barrel was represented exactly to dimension, as a 23.5” long tube

with inner and outer diameter’s to match the actual rifle barrel geometry. A pre-stressed Modal

Analysis method was used to simulate a force moving down the barrel. The receiver end of the

barrel was fixed by constraint. The barrel itself was constrained with a remote displacement

to allow only displacement in the Y axis (vertical component), and rotation of each degree of

freedom about the X axis. Z axis deflection and oscillation were considered negligible for this

analysis.

Using the Static Structural Module to create the pre-stressed environment, a force was applied

to each discretized segment of the rifle barrel in steps, with the force moving to the next segment

with each step increment. The Static Structural analysis was then fed into a Modal Analysis

Module for the final modal calculations of modeshape and natural frequencies.

The rifle barrel was set up in ANSYS as:

Figure 25: ANSYS Static Structural Setup

With constraints applied, the moving force function was arranged to shift to the next beam

segment with each step of the iteration:


4 ANALYTICAL MODELING RESULTS USING ANSYS 29

Figure 26: ANSYS Static Structural Moving Force, Step 1

The Modal Analysis was set up to calculate modal displacement. The magnitude of deflection

is not true to the actual system; it is scalable by any factor, similar to the modeshapes matrix

ψ. Modes 1 through 3 are shown below.

Figure 27: ANSYS Modal Analysis of Tikka T3X System, Mode 1, 38.6 Hz
4 ANALYTICAL MODELING RESULTS USING ANSYS 30

Figure 28: ANSYS Modal Analysis of Tikka T3X System, Mode 2, 240.7 Hz

Figure 29: ANSYS Modal Analysis of Tikka T3X System, Mode 3, 669.8 Hz

The ANSYS Modal Analysis for the first three modes of the Tikka T3X system resulted in nat-

ural frequencies of 38.6, 240.7, and 669.8 Hz. Modeshapes matched the discretized mathematical

model within 7%, and also the Blevins calculations(8) to within 3%.
5 EMPIRICAL RESULTS 31

5 Empirical Results

Before proposing any changes to the Tikka system, especially as an improvement, it is important

to correlate the models developed with real world empirical data. To that end, a Tikka T3X rifle

was utilized to collect live fire data with a DEWESoft Sirius (10) Data Acquisition (DAQ) System

and PCB Accelerometers. It proved to be difficult to collect valid data due to high acceleration

loading during firing events, and resultant low data resolution which prompted a second live fire

event. The data and conclusions presented below are arranged in the order easiest for the reader

to follow, and not in the order in which the data were collected.

Testing was set up at a private range facility, and utilized twin +/-500G accelerometers

mounted to the rifle barrel. The action time for primer strike to projectile exit was assessed as

occurring within 2 milliseconds, which drove the choice for channel sample rate. This equated
1 1
to an event frequency of actiontime = .002 = 500Hz. In terms of signal processing, the Nyquist

Theorem dictates that, in order to capture frequency content with known error in peak magnitude,

we must sample at least twice the frequency of event occurrence. Sampling faster than 2x

increases the accuracy of the peak magnitudes, for example, at the dominant frequencies. The

more oversampled the data set, the more accurate the frequency content magnitudes, and the

smaller each frequency bin may be (x axis resolution) for a frequency domain transform. Common

industry measurement practice is to oversample by a factor of at least 10.


5 EMPIRICAL RESULTS 32

Figure 30: Tikka System Rifle Barrel Test, DEWESoft DAQ, PCB Accels

In order to understand free vibration modes and magnitudes, a tap test was performed on the

rifle barrel. Accelerometers were placed at the muzzle and at 5 inches from the muzzle towards

the breech end of the barrel. The accelerometers were secured to the barrel using hose clamps;

one placed at the muzzle end, and one placed 5 inches back towards the breech end from the

muzzle. A sample rate of 100kHz was selected for the tap test to ensure data resolution during

post processing. The DAQ system was set up to apply an antialiasing Butterworth filter at

50kHz. Higher order filters are normally used when sharp data attenuation is desirable. Figure

31 demonstrates a Butterworth filter response dependant on the utilized filter order.


5 EMPIRICAL RESULTS 33

Figure 31: Butterworth Filter Response Depending on Filter Order

(20)

With the system live and collecting data, the barrel was rapped sharply with a steel screw-

driver to excite its natural modes.

Figure 32: Tikka System Rifle Barrel Test, Tap Test, 100kHz
5 EMPIRICAL RESULTS 34

The tap test results were analyzed within a 0.002 second window of time, and cleanly show the

impulse, free vibration of the system, and the modes of oscillation in the Fast Fourier Transform

(FFT). Data was plotted up to 500 Hz as the window of interest since high frequency modes

were orders of magnitude lower in amplitude. Note that the first and second barrel modes are

dominant in the FFT, along with other modes excited within the system. This is expected since

a complete rifle system is the subject of this test, not just the rifle barrel.

The first live fire test mimicked the set up for the tap test, accelerometers were placed at the

muzzle end and at approximately 3 inches aft from the muzzle end towards the breech end. Ac-

celerometer rigid wax was used to fix the accelerometers in place, and held consistent throughout

each round fired, although lower magnitude accelerations resulted in this measurement iteration.

This is assessed as due to the coupling between the barrel and accelerometers; the accelerometer

wax was not as stiff as the hose clamps used for the tap test and second live fire test. The rifle

itself was held rigidly by a Pig Saddle and aluminum field shooting tripod from Shadow Tech(9).

The rifle was man fired for each data set collected. Again, the DEWESoft DAQ system was set

up for two channel data collection, this time at 20kHz per channel, with an antialiasing filter

placed at 8kHz. The resultant data is shown below.

Figure 33: Tikka System Rifle Barrel Test, Accelerometer Placement


5 EMPIRICAL RESULTS 35

Figure 34: Tikka System Rifle Barrel Test, Time Series Acceleration Data, Test 1, Shot 1

Accelerometer data is windowed from the moment of sear release (the moment the trigger

is pressed) to 0.002 seconds later, this being the action time of the system. Data resolution is

immediately apparent since only approximately 38 data points are available for review. Acceler-

ation magnitudes are within the accelerometer capability. The accelerometer channels were then

integrated twice to obtain displacement. The actual integration algorithm was managed internal

to the DEWESoft platform, and was high pass filtered below 0.1Hz to reduce drift and error.

Total barrel displacement for the action time window is 0.007 for the muzzle and 0.002 for the

mid barrel accel.

The time series acceleration data was then transformed to the frequency domain by means of

an FFT within DEWESoft. Filtering was implemented according to DEWESoft best practices

and the author’s prior data collection knowledge. That is to say, a band pass filter was imple-

mented from .1 Hz to 10 kHz to attenuate the transform DC offset at low frequency and high

frequency noise.

Figure 35 below displays the empirically determined modes of oscillation for the Tikka T3X

system during a live fire event.


5 EMPIRICAL RESULTS 36

Figure 35: Tikka System Rifle Barrel Test, Empirically Determined Modes, Test 1, Shot 1

Shot 2 was performed in exactly the same manner, and the subsequent data is shown below

in Figures 36 and 37. Again note that the time window was 0.002 seconds from sear release.

Figure 36: Tikka System Rifle Barrel Test, Time Series Acceleration Data, Test 1, Shot 2
5 EMPIRICAL RESULTS 37

Figure 37: Tikka System Rifle Barrel Test, Empirically Determined Modes, Test 1, Shot 2

The modes from shot 2 are consistent with that from shot 1, with both the first and second

barrel mode being prevalent at 32.3Hz and 263.7Hz.

After reviewing the live fire data from test one, the determination was made that the data were

insufficient to proceed, primarily based on the resolution of the frequency content. The assessment

was that the frequency data did not have enough frequency bins for the x axis to adequately

determine resonant modes and peaks. It was also desirable to have higher resolution data to

reduce error for the double integration to obtain displacement. To that end, the DEWESoft

DAQ system was again secured and set up again for a two channel system (vertical barrel motion

only), sampled at 100 kHz per channel. In this way, the data were amply oversampled to assure

decent frequency content and magnitude for later use in system optimization.

Setup and accelerometer filters were adjusted slightly to optimize the Butterworth bandpass

filter applied for anti-aliasing from .8 Hz to 30 kHz. The filter was implemented using a 6th

order algorithm to accentuate the filter rolloff and data attenuation closer to the filter edges.

The accelerometers were fixed to the barrel using hose clamps for greater rigidity and coupling.

Accel 1 was located at the muzzle end, and Accel 2 was at -5 inches from the muzzle back towards

the breech end.


5 EMPIRICAL RESULTS 38

Two shots were expended for test 2, and even with +/- 500G accelerometers, the sensors

experienced overload errors and railed at measured magnitudes (the sensor signal input went

beyond the capability of the sensor), and the resultant frequency content was erroneous. Upon

further investigation, the accelerometers measured accurately during the action time of the sys-

tem, but grossly exceeded their capability once the projectile left the barrel and the totality

of the impulse response played out. A specific example of this is displayed below in Figure 38.

Note the accelerometer time series magnitude offset with corresponding large y axis value outside

the calibrated range of the device. The frequency versus acceleration plot shows low frequency

content with a progressive attenuation down to zero magnitude, in no way corresponding to the

impulse that the system experienced during the firing event. While interesting, this data is not

usable for this analysis except to help select higher range accelerometers for future testing.

Figure 38: Saturated Accelerometer, Test 2, Shot 2


5 EMPIRICAL RESULTS 39

The author asserts that the acceleration data within the action time of the system is still

valid, but the follow on accel data is garbage and cannot be used. The double integration of

the signals in shot 2 used for displacement resulted in values that increased to a nonsensical

order of magnitude before dropping off again; clearly laden with error related to saturating the

accelerometer’s measurement capability. Additionally, a frequency transform of the accelerometer

time series was error fraught, with extensive drift and offset. See Figures 39 and 40; the asterisk

notes test samples that are suspect due to instrumentation error.

Figure 39: Tikka System Rifle Barrel Test, Time Series Acceleration Data, Test 2, Shot 1
5 EMPIRICAL RESULTS 40

Figure 40: Tikka System Rifle Barrel Test, Time Series Acceleration Data, Test 2, Shot 2

Table 1: Displacements from Empirical Test Results (* denotes data in question)


Test Shot Muzzle Disp (in) Mid Barrel Disp(in) Muzzle Angle (deg)

1 1 .007 .002 .122

1 2 .006 .001 .122

2 1 .001* .020* .22*

2 2 .030* .030* 0*

Muzzle angle is derived from the measured slope between the max displacement of each

accelerometer. In other words, it is the slope between two points of the modeshape at the end

of the barrel. Note that the error at 100 yards, or 3,600 inches of distance equates to a vertical

delta of 7.67 inches for a muzzle angle of 0.122 degrees, and half that for the muzzle angle of 0.10

degrees. Obviously, this error compounds at distance. Normal midwest shots for hunting big

game exceed 600 yards distance, which equates to an 46 inch error at the target, solely due to

muzzle angle and displacement. In military engagements, marksmen may engage targets beyond

1,500 yards, which in this case equates to a 115 inch error at the target independent of nominal
6 BARREL OPTIMIZATIONS 41

projectile drop and wind calls. If an average human is 20 inches across at the shoulders and 6

feet tall, this error is obviously significant and will result in a target miss without correction.

From the data collected, several items are apparent. The first is that there are modes from

the entire system oscillating besides just the barrel modes. As mentioned prior, this is expected

since the complete rifle is fitted with an optic and is shooting from a rigid tripod mount. It is

however, obvious that the first two barrel modes predominantly impact system oscillations, and

in some of the empirical test cases, dominate the vibration energy content. Second is that the

empirical displacement data aligns with the first barrel modeshape at 36 Hz for all correlated

shots. For shots where the accelerometers exceeded their magnitude capability or exhibited

response errors, that data was discounted. Last is that the tap test data validates the frequencies

and modeshapes of the analytical (FEA) and mathematical models to within 10%. Moving

forward, the mathematical and analytical models will be used to suggest optimizations for the

rifle system, predominately targeting the first and second barrel modes.

6 Barrel Optimizations

There are several opportunities for altering and/or optimizing barrel vibrations. They will be

presented here in order of complexity. In this case, barrel optimization equates to better precision,

or less error at the target distance, which is directly tied to muzzle tip displacement and angle

of rotation at projectile exit. Both ANSYS and the MATLAB mathematical model are used to

demonstrate the various aspects of barrel improvements. In the case of the MATLAB model, the

data is normalized to the Test 1 data magnitudes. Using this algorithm, the stock Tikka T3X

system characteristics, at projectile exit, are:

Table 2: Tikka T3X Baseline Muzzle Performance at Projectile Exit


Mode Muzzle Disp (in) Projo Exit Angle (deg) 100 yard Vertical Error(in)

1 .0016 .0055 .3354

2 .0053 -.0768 -4.8265


6 BARREL OPTIMIZATIONS 42

6.1 Barrel Length

An easy and obvious way to improve barrel deflection during firing events is to make the barrel

shorter. From the beam shape equations (6), we know that barrel length exponentially affects

barrel stiffness via the reaction force and moment equations:

12EI
F = (38)
L3

6EI
F = (39)
L2

By lowering the overall barrel length, and thus exponentially reducing the denominator of the

reaction forces equation, we achieve a stiffer barrel as defined by the new stiffness matrix [K].

In the example of the Tikka T3X system, the mathematical model predicts first and second

modes at 35.7 and 224.8 Hz for the 23.5 inch length barrel. Mode 1 is predicted to cause .0016

inches of deflection at the muzzle, and mode two will cause .0055 inches of deflection at the

muzzle. The predicted angular error at the muzzle tip from the second mode is .077 degrees,

which equates to a vertical error of 4.83 inches at 100 yards or 48.3 inches at 1000 yards.

Now, if the barrel length is adjusted to one half of that length, we see an immediate improve-

ment in the system’s resonant modes. Mode 1 is shifted from 35.7 Hz to 143 Hz, and Mode 2 is

shifted from 224.8 Hz to 899 Hz. Muzzle displacement drops by approximately 60% from .0055

inches to .0021 inches when excited at Mode 2. And it follows that the shorter, stiffer barrel,

likewise has a lower muzzle tip angle at the cusp of projectile exit resulting in lower targeting

error at distance.

Table 3: Tikka T3X System L/2 Barrel Performance


Mode Muzzle Disp (in) Projo Exit Angle (deg) 100 yard Vertical Error(in)

1 .0006 .004 .2498

2 .0021 -.0588 -3.6967


6 BARREL OPTIMIZATIONS 43

Figure 41: Short Barrel Optimization vs. Tikka T3X System Response
6 BARREL OPTIMIZATIONS 44

Figure 42: Short Barrel Optimization vs. Tikka T3X System Action Time Response

This seems like an easy fix; why not make all rifle barrels shorter? The unfortunate trade

off is that interior ballistics demands a certain barrel length to achieve powder burn efficiency,

pressure curves, length of barrel to stabilize the projectile, and a muzzle velocity that is not

severely impacted as to prevent long distance shooting in the first place. From the end user’s

perspective, shorter barrels are not always bad, but lower muzzle velocity and thus terminal

energy at the target range, equate to less lethality, projectile stability (supersonic, transonic, and

subsonic), and overall accuracy. Barrel lengths must be chosen carefully to optimize not only

barrel harmonics, but also muzzle velocity, projectile performance, and overall system weight.

Figure 43 shows two industry prototype options that the author has successfully shot out to 1100

yards with standard ammunition, demonstrating that long barrels are not always necessary to
6 BARREL OPTIMIZATIONS 45

shoot precisely.

Figure 43: 12.5” Mid Gas System in 5.56mm (L) and 12.5” Piston System in 5.56mm (R)

6.2 Barrel Moment of Inertia

Another method to optimizing barrel performance is to increase its Area Moment of Inertia, or

variable I used in the beam stiffness equations. Adjustments to I impact the overall stiffness

matrix [K] as part of the equations of motion, and subsequent reaction loads for each degree of

freedom; the higher the Area Moment of Inertia, the greater the beam stiffness, the lower the

beam deflection for each unit of force applied.

Manufacturers make barrel profiles currently that are more complex than simply round or

tapered cylinders. Profiles include hexagonal, diamond, and fluted. Fluting is marketed as an

improvement for both barrel stiffness and heat transfer, although improvement of these claims is

dubious at best. There is also the example cited earlier in this paper regarding the structured rifle

barrels and their claim of being ”harmonically dead”. These are all examples of manufacturers
6 BARREL OPTIMIZATIONS 46

attempting to optimize barrel performance by increasing the Area Moment of Inertia.

The optimization studied in this paper is an ”H” shaped cross section extrusion. From a

vibrations perspective, the Area Moment of Inertia is 2.9 times greater than the stock barrel in

the Tikka T3X system, and length and material modulus are exactly the same. This increase

in Area Moment of Inertia shifts the First Mode to 51.5 Hz, and the Second Mode to 319.8 Hz.

The following improvements to muzzle displacement, projectile exit angle, and vertical error at

100 yards are listed in Table 4.

Table 4: Optimization of Grid Style Extruded Barrel


Mode Muzzle Disp (in) Projo Exit Angle (deg) 100 yard Vertical Error(in)

1 .0001 .004 .026

2 .0005 .007 .429

Figure 44: Grid Style Extruded Barrel Blank Model, Area Moment of Inertia Optimization
6 BARREL OPTIMIZATIONS 47

Figure 45: Grid Style Extruded Barrel Blank, Modal Analysis, Mode 2

Inputing the moment of inertia improvement into the mathematical model, the response

comparison may be made as shown below. Note that Figure 46 shows total system response and

Figure 47 shows only the action time of the cartridge and system combination.
6 BARREL OPTIMIZATIONS 48

Figure 46: Moment of Inertia Optimization vs. Tikka T3X System Response
6 BARREL OPTIMIZATIONS 49

Figure 47: Moment of Inertia Optimization vs. Tikka T3X System Action Time Response

The improvements in muzzle displacement and angle at projectile exit represent an order of

magnitude improvement, which equates to significantly less vertical error at the target. The

geometry maximizes stiffness in bending and also lowers weight as compared to the nominal

rifle barrel. The trade off in this case is complexity in manufacturing. The barrel blank itself

may be extruded or forged on a mass production scale using standard manufacturing techniques.

However, follow on machining to fit the barrel to the receiver would likely be cumbersome.

Depending on the desired caliber of manufacture, and peak pressure of the round, the barrel

would have to gain cross sectional thickness to support the necessary loads at operating pressures,

which could be a detractor for weight eventually. The barrel blank rifling could be gun drilled

like a standard legacy rifle barrel, but not cold hammer forged, which is a much more common
6 BARREL OPTIMIZATIONS 50

commercial practice today. Still, it presents some interesting improvements strictly from an

overall barrel performance perspective.

6.3 Modal Mass Tuning

A third improvement for rifle barrel vibration and precision is the concept of Modal Mass Tuning.

The Modal Mass matrix, or variable [µ] in the mathematical algorithm used to assess these

optimizations, is used in the calculations for system response. As stated above, the Modal Mass

matrix derived from the transformation equation is:

T
[µ] = {ψ} [M ] {ψ} (40)

This results in mass applied to each degree of freedom proportional to the modeshape at

each mode. Thus, [µ] is not simply the total mass of the system divided by the displacement

degrees of freedom. It is a diagonal matrix, but the values of µii are not all equal. It may be

shown using the mathematical algorithm for this system’s response, that adding mass to the last

two degrees of freedom improve precision by reducing muzzle displacement and muzzle angular

error at projectile exit. However, if a tuning system were installed on the rifle that allowed

placement of masses at the longitudinal locations of the discretized masses of the barrel, it then

becomes possible to tune the response to something more favorable. For example, if the vector

ζ = [1, 1, 1, 5, 1, 5, 1, 5, 1, 5] were multiplied by the mass matrix M , the response of the system

entirely changes. This operation mathematically represents tuning masses placed at nodes 10, 8,

6, and 4 that sum to five times the discretized modal mass at those locations i.e. a series of masses

precisely mounted at nodal points along the vibrating beam. The impact in the mathematical

system is a significant improvement in both muzzle displacement and muzzle tip angle at the

time of projectile exit.


6 BARREL OPTIMIZATIONS 51

Table 5: Tikka T3X System Tuned Nodal Masses Performance


Mode Muzzle Disp (in) Projo Exit Angle (deg) 100 yard Vertical Error(in)

1 .0002 .0006 .3550

2 .0006 -0088 -.5500

The Modal Mass Tuning method of system optimization response may be compared to the

nominal response in the mathematical model easily. Using the parameters above, the following

plots were generated; the first for total system response through 0.1 seconds, and the second only

plotted through the action time of the cartridge and system being fired.

Figure 48: Modal Mass Tuning Scenario vs Nominal Tikka T3X Vibration Response
7 CONCLUSIONS 52

Figure 49: Grid Style Extruded Barrel Blank, Modal Analysis, Mode 2

As compared to the baseline performance in Table 2, modal mass tuning using this specific

mass arrangement, accounts for an 89% vertical error reduction from the mode 2 offset at 100

yards.

7 Conclusions

Regarding the topic of rifle barrel optimizations for muzzle tip displacement and projectile exit

angle, there are several optimizations that are possible. Within this paper, three distinct methods

of altering barrel response from the nominally offered product were reviewed. A rather simple

optimization, if the system and end user can tolerate reduction in muzzle velocity, is to reduce

the overall length of the barrel. By changing barrel length, the overall barrel stiffness is increased,

the resonant modes are shifted higher, and the muzzle tip displacement and angle at the moment
7 CONCLUSIONS 53

of projectile exit are favorable when compared to the nominal barrel length. For users who desire

a given barrel length for terminal performance, exploring a barrel cross sectional geometry that

provides an increased moment of inertia (2nd area moment), provides exceptional improvement to

the barrel dynamics. Specifically, the ”H” profile analyzed within this thesis offers very favorable

improvement in both muzzle tip displacement and projectile exit angle, which translates to

the reduction of error at the desired target range. In this case, users trade these favorable

characteristics with a potential increase in overall system weight (dependent on necessary barrel

wall thickness) and increased cost of manufacturing. Lastly, the concept of a tuning mass system

was explored, where masses were placed at nodes 10, 8, 6, and 4 to tune the barrel response.

This capability to tune a rifle barrel may be commercially adopted by the addition of sliding

masses on the barrel, with set screws. Alternately, a gear profile cut into a rib on the top of the

barrel could be used to mechanically increment masses longitudinally along the barrel to achieve

the same desired end effect. Analytical results of this technique demonstrated very favorable

response and muzzle angle results. Further, this type of system could be tuned to a particular

projectile mass and powder charge, increasing the potential for precision and reduction of error

at the target. The trade off is manufacturing cost, and overall complexity of such a system. It

is unlikely that such a system would be used by your average sportsman or military member.

Rather, it seems like a system for the bench rest shooter or precision rifle user.

In any case, these solutions explored are all very interesting, and my be desirable to the

commercial and military developers of precision firearm systems and high quality barrel suppliers.
REFERENCES 54

References
[1] Tikka Products, www.tikka.fi, 2022
[2] https://www.hunter-ed.com, Kalkomey, 2022
[3] http://ffden-2.phys.uaf.edu, M. Tilly, 2019
[4] https://www.luckygunner.com, Andrew, 2012
[5] ”Ballistics: Theory and Design of Guns and Ammunition”, 2nd Edition, D. Carlucci, S.
Jacobsen, 2013
[6] ”Structural Dynamics Theory and Application”, J. Tedesco, W. McDougal, C. A. Ross, 1999
[7] ”The Dynamical Behaviour of Structures”, G. B. Warburton, 1964
[8] ”Formulas for Natural Frequency and Modeshape”, R. Blevins, 2001
[9] https://hogsaddle.com/, Shadow Tech, 2022
[10] https://dewesoft.com/products/daq-systems/sirius, DEWESOFT, 2022
[11] https://training.dewesoft.com/online/course/, Sound Level Frequency Weighting Curves,
DEWESOFT, 2022
[12] ”Vibrations of Rifle Barrels”, A. Mallock, May 2, 1901
[13] ”Study on Barrel Vibration Characteristics of Typical Sniper Rifle”, Xu, Guan, Liu, Xu,
2018
[14] ”Concept of a Gun Barrel Based on the Layer Composite Reinforced with Continuous Fila-
ment”, Pyka et al., 2019
[15] ”Analysis and Control of Gun Barrel Vibrations”, F. Buyukcivelek, 2011
[16] ”A Review on the Gun Barrel Vibrations and Control for a Main Battle Tank”, Dursun,
Buyukcivelek, Utlu, 2017
[17] ”Dynamic Analysis of Gun Barrel Vibrations Due to Effect of an Unbalanced Projectile Con-
sidering 2-D Transverse Displacements of Barrel Tip Using a 3-D Element Technique”,
Koc, Esen, Cay, 2018
[18] ”Using Barrel Vibrations to Tune a Barrel”, G. Kolbe, 2021
[19] https://tacomhq.com/structured-barrels, 2021
[20] https://www.electronicshub.org/butterworth-filter, 2021
[21] ”Ammunition, Demystified. The (non) Bubba’s Guide to How Ammo Really Works”, J.
Siewert, 2022
[22] https://www.shootingsoftware.com/barrel.htm, 2004
[23] https://www.shootinguk.co.uk/guns/gun-maintenance/rifle-accuracy-87614, 2020
8 APPENDIX 55

8 Appendix

8.1 Barrel Vibrations MATLAB Code

%this program calculates the stiffness matrix for a cantilever beam with 20

%degrees of freedom.

clear all

clc

close all

%barrel geometry in inches

% OD = input(’What is the barrel Outer Diameter in inches?’);

% ID = input(’What is the barrel Inner Diameter in inches?’);

% L = input(’What is the barrel overall length in inches?’);

OD=.75; % in inches

ID=.22; % in inches

%____________________________________

at = .002; %action time in seconds

L = 23.5; %barrel length in inches

%____________________________________

n = 10; %number of discretized nodes

r1 = OD/2;

r2 = ID/2;

%system values

den = .285; %lbs/in^3

A = pi*(r1^2 - r2^2); %in^2

W = den*L*A; %lbs

m = W/386.4; % lbs*sec^2/in

E = 29e6; %modulus in lbs/in^2%

I = (pi/64)*(OD^4 - ID^4); %in^4

%I = 3.2;

l = L/n; %in

U = 11750; %velocity of projectile in in/sec

dt=0.0001;

tmax=.1;

t=0:dt:tmax;

nsteps=length(t);

% %load influence matrix

r=1+zeros(10,1);

%mass matrix nDOF

for j=1:n

M(j) = m/n;

end
8 APPENDIX 56

M = diag(M);

% tune = [1 1 1 5 1 5 1 5 1 5];

% M = M*diag(tune);

%Generate Stiffness Matrix

%DOF1

k1=[24*E*I/l^3;-12*E*I/l^3;0;0;0;0;0;0;0;0;0;6*E*I/l^2;0;0;0;0;0;0;0;0];

%DOF2

k2=[-12*E*I/l^3;24*E*I/l^3;-12*E*I/l^3;0;0;0;0;0;0;0;-6*E*I/l^2;0;6*E*I/l^2;0;0;0;0;0;0;0];

%DOF3

k3=[0;-12*E*I/l^3;24*E*I/l^3;-12*E*I/l^3;0;0;0;0;0;0;0;-6*E*I/l^2;0;6*E*I/l^2;0;0;0;0;0;0];

%DOF4

k4=[0;0;-12*E*I/l^3;24*E*I/l^3;-12*E*I/l^3;0;0;0;0;0;0;0;-6*E*I/l^2;0;6*E*I/l^2;0;0;0;0;0];

%DOF5

k5=[0;0;0;-12*E*I/l^3;24*E*I/l^3;-12*E*I/l^3;0;0;0;0;0;0;0;-6*E*I/l^2;0;6*E*I/l^2;0;0;0;0];

%DOF6

k6=[0;0;0;0;-12*E*I/l^3;24*E*I/l^3;-12*E*I/l^3;0;0;0;0;0;0;0;-6*E*I/l^2;0;6*E*I/l^2;0;0;0];

%DOF7

k7=[0;0;0;0;0;-12*E*I/l^3;24*E*I/l^3;-12*E*I/l^3;0;0;0;0;0;0;0;-6*E*I/l^2;0;6*E*I/l^2;0;0];

%DOF8

k8=[0;0;0;0;0;0;-12*E*I/l^3;24*E*I/l^3;-12*E*I/l^3;0;0;0;0;0;0;0;-6*E*I/l^2;0;6*E*I/l^2;0];

%DOF9

k9=[0;0;0;0;0;0;0;-12*E*I/l^3;24*E*I/l^3;-12*E*I/l^3;0;0;0;0;0;0;0;-6*E*I/l^2;0;6*E*I/l^2];

%DOF10

k10=[0;0;0;0;0;0;0;0;-12*E*I/l^3;12*E*I/l^3;0;0;0;0;0;0;0;0;-6*E*I/l^2;-6*E*I/l^2];

%DOF11

k11=[0;-6*E*I/l^2;0;0;0;0;0;0;0;0;8*E*I/l;2*E*I/l;0;0;0;0;0;0;0;0];

%DOF12

k12=[6*E*I/l^2;0;-6*E*I/l^2;0;0;0;0;0;0;0;2*E*I/l;8*E*I/l;2*E*I/l;0;0;0;0;0;0;0];

%DOF13

k13=[0;6*E*I/l^2;0;-6*E*I/l^2;0;0;0;0;0;0;0;2*E*I/l;8*E*I/l;2*E*I/l;0;0;0;0;0;0];

%DOF14

k14=[0;0;6*E*I/l^2;0;-6*E*I/l^2;0;0;0;0;0;0;0;2*E*I/l;8*E*I/l;2*E*I/l;0;0;0;0;0];

%DOF15

k15=[0;0;0;6*E*I/l^2;0;-6*E*I/l^2;0;0;0;0;0;0;0;2*E*I/l;8*E*I/l;2*E*I/l;0;0;0;0];

%DOF16

k16=[0;0;0;0;6*E*I/l^2;0;-6*E*I/l^2;0;0;0;0;0;0;0;2*E*I/l;8*E*I/l;2*E*I/l;0;0;0];
8 APPENDIX 57

%DOF17

k17=[0;0;0;0;0;6*E*I/l^2;0;-6*E*I/l^2;0;0;0;0;0;0;0;2*E*I/l;8*E*I/l;2*E*I/l;0;0];

%DOF18

k18=[0;0;0;0;0;0;6*E*I/l^2;0;-6*E*I/l^2;0;0;0;0;0;0;0;2*E*I/l;8*E*I/l;2*E*I/l;0];

%DOF19

k19=[0;0;0;0;0;0;0;6*E*I/l^2;0;-6*E*I/l^2;0;0;0;0;0;0;0;2*E*I/l;8*E*I/l;2*E*I/l];

%DOF20

k20=[0;0;0;0;0;0;0;0;6*E*I/l^2;-6*E*I/l^2;0;0;0;0;0;0;0;0;2*E*I/l;4*E*I/l];

K = [k1 k2 k3 k4 k5 k6 k7 k8 k9 k10 k11 k12 k13 k14 k15 k16 k17 k18 k19 k20];

%symmetry check on stiffness matrix K

tf = issymmetric(K);

if tf == 1;

waitfor(msgbox(’K is symmetric’,’Success’));

else

waitfor(msgbox(’K has an error in it (not symmetric)’,’Error’,’error’));

return;

end

%Static condensation of the stiffness matrix

Ktt = K(1:10,1:10);

Kt0 = K(1:10,11:20);

K0t = K(11:20,1:10);

K00 = K(11:20,11:20);

Kt = Ktt - Kt0*inv(K00)*K0t; %static condensation of the stiffness matrix

[sai,lambda] = eig(Kt,M)

omega = sqrt(lambda);

freq = (omega/(2*pi))

%____________________________________________________________________

%modeshape variables for cantilever beams |

%____________________________________________________________________|

Rho = [1.87510407 4.69409113 7.85475744 10.99554073 14.13716839];

sigma = [.734095514 1.018467319 .999224497 1.000033553 .999998550];

%calculate modal influence factor for first three modes ref: Blevins

%_____________________________________________________________________

%Dynamic Loading

P=1; %moving load magnitude*****************************************

MM = transpose(sai)*M*sai; %modal mass matrix


8 APPENDIX 58

counter = 1;

for j=1:nsteps

if t(j)<= at

MSfac1 = (cosh(Rho(1)*U*t(j)/L) - cos(Rho(1)*U*t(j)/L) - sigma(1)*(sinh(Rho(1)*U*t(j)/L) - sin(Rho(1)*U*t(j)/L)));

MSfac2 = (cosh(Rho(2)*U*t(j)/L) - cos(Rho(2)*U*t(j)/L) - sigma(2)*(sinh(Rho(2)*U*t(j)/L) - sin(Rho(2)*U*t(j)/L)));

MSfac3 = (cosh(Rho(3)*U*t(j)/L) - cos(Rho(3)*U*t(j)/L) - sigma(3)*(sinh(Rho(3)*U*t(j)/L) - sin(Rho(3)*U*t(j)/L)));

b = .04; % normalization factor

phi_a1(j) = b*MSfac1; %Warburton eq 4.7

F_a1(j) = (P*phi_a1(j))/MM(10,10); %Warburton eq 3.30

phi_a2(j) = b*MSfac2;

F_a2(j) = (P*phi_a2(j))/MM(9,9);

phi_a3(j) = b*MSfac3;

F_a3(j) = (P*phi_a3(j))/MM(8,8);

counter = counter +1;

else

phi_a1(j) = 0;

F_a1(j) = 0;

phi_a2(j) = 0;

F_a2(j) = 0;

phi_a3(j) = 0;

F_a3(j) = 0;

end

end

%response calculations

damp=(.05+zeros(10,1)); %damping at 5%

xi=damp;

omegad=sqrt(1-xi.^2).*omega’;

%response of first mode

%h1=(1/(omegad(10,10)))*exp(-omega(10,10)*xi(1)*t).*sin(omegad(10,10)*t);

h1=(1/(omegad(10,10)*MM(10,10)))*exp(-omega(10,10)*xi(1)*t).*sin(omegad(10,10)*t);

z1=dt*conv(h1,sai(:,10)’*r*F_a1);

q1=sai(:,10)*z1;

%response of second mode

%h2=(1/(omegad(9,9)))*exp(-omega(9,9)*xi(2)*t).*sin(omegad(9,9)*t);

h2=(1/(omegad(9,9)*MM(9,9)))*exp(-omega(9,9)*xi(2)*t).*sin(omegad(9,9)*t);

z2=dt*conv(h2,sai(:,9)’*r*F_a2);
8 APPENDIX 59

q2=sai(:,9)*z2;

%response of third mode

%h3=(1/(omegad(8,8)))*exp(-omega(8,8)*xi(3)*t).*sin(omegad(8,8)*t);

h3=(1/(omegad(8,8)*MM(8,8)))*exp(-omega(8,8)*xi(3)*t).*sin(omegad(8,8)*t);

z3=dt*conv(h3,sai(:,8)’*r*F_a3);

q3=sai(:,8)*z3;

%total response

x(1,:)=q1(1,:)+q2(1,:)+q3(1,:);

x(2,:)=q1(2,:)+q2(2,:)+q3(2,:);

x(3,:)=q1(3,:)+q2(3,:)+q3(3,:);

%plotting------------------------------------------------------------

figure

plot(t,x(1,1:nsteps),’k’,’LineWidth’,2)

hold

plot(t,q1(1,1:nsteps),’b’)

plot(t,q2(1,1:nsteps),’r’)

plot(t,q3(1,1:nsteps),’g’)

legend(’Total Response’,’Mode 1’,’Mode 2’,’Mode 3’)

title(’Total Response at First Mode’)

figure

plot(t,x(2,1:nsteps),’k’,’LineWidth’,2)

hold

plot(t,q1(2,1:nsteps),’b’)

plot(t,q2(2,1:nsteps),’r’)

plot(t,q3(2,1:nsteps),’g’)

legend(’Total Response’,’Mode 1’,’Mode 2’,’Mode 3’)

title(’Total Response at Second Mode’)

% figure

% plot(t,x(3,1:nsteps),’k’,’LineWidth’,2)

% hold

% plot(t,q1(3,1:nsteps),’b’)

% plot(t,q2(3,1:nsteps),’r’)

% plot(t,q3(3,1:nsteps),’g’)

% legend(’Total Response’,’Mode 1’,’Mode 2’,’Mode 3’)

% title(’Total Response at Third Mode’)

ratios = [freq(10,10)/freq(10,10) freq(9,9)/freq(10,10) freq(8,8)/freq(10,10) freq(7,7)/freq(10,10)];

%calculate slopes between each node

%mode 1

m1node9 = sai(9,10);

m1node10 = sai(10,10);

pkdisp1 = max(x(1,1:counter))

m1slope = ((m1node10-m1node9)*pkdisp1)/(L/n);
8 APPENDIX 60

%mode 2

m2node9 = sai(9,9);

m2node10 = sai(10,9);

pkdisp2 = max(x(2,1:counter))

m2slope = ((m2node10-m2node9)*pkdisp2)/(L/n);

%mode 3

m3node9 = sai(9,8);

m3node10 = sai(10,8);

pkdisp3 = max(x(3,1:counter))

m3slope =((m3node10-m3node9)*pkdisp3)/(L/n);

%angle of barrel rotation at muzzle

angle_m1 = atand(m1slope)

angle_m2 = atand(m2slope)

angle_m3 = atand(m3slope)

%calculate deviation at 100 yards

rng = 100*3*12;%range in inches

peak_disp_muzzle = [1 2 3; pkdisp1 pkdisp2 pkdisp3]

angles = [1 2 3; angle_m1 angle_m2 angle_m3]

v_error_100 = [rng*tand(angle_m1),rng*tand(angle_m2), rng*tand(angle_m3)]%vertical deviation in inches

resp1_out = transpose([t(1:nsteps); x(1,1:nsteps)]);

resp2_out = transpose([t(1:nsteps); x(2,1:nsteps)]);

8.2 MATLAB Code Output Example

sai =

0.6053 0.9645 -1.0000 1.0000 -0.9153 0.7315 0.4866 0.2805 0.1092 0.0163

-0.7825 -1.0000 0.6101 -0.0397 -0.5761 0.9835 1.0000 0.7625 0.3590 0.0624

0.9282 0.7768 0.1347 -0.8637 0.8826 -0.1470 0.6792 1.0000 0.6379 0.1336

-1.0000 -0.3053 -0.7928 0.7293 0.3804 -1.0000 -0.2687 0.7782 0.8492 0.2257

0.9948 -0.2614 0.8919 0.2962 -0.9603 -0.2581 -0.9234 0.1786 0.9215 0.3343

-0.9128 0.7462 -0.3632 -0.9603 -0.1977 0.8968 -0.6374 -0.4882 0.8169 0.4555

0.7604 -0.9971 -0.4210 0.4547 1.0000 0.6385 0.2855 -0.8579 0.5337 0.5856

-0.5496 0.9365 0.9110 0.6102 0.0179 -0.6081 0.8826 -0.7068 0.1021 0.7213

0.2916 -0.5709 -0.7397 -0.8958 -0.9446 -0.8059 0.4781 -0.0584 -0.4273 0.8601
8 APPENDIX 61

-0.0758 0.1586 0.2321 0.3394 0.4721 0.6077 -0.7163 0.8638 -1.0000 1.0000

lambda =

1.0e+09 *

2.1880 0 0 0 0 0 0 0 0 0

0 1.7351 0 0 0 0 0 0 0 0

0 0 1.1866 0 0 0 0 0 0 0

0 0 0 0.7097 0 0 0 0 0 0

0 0 0 0 0.3722 0 0 0 0 0

0 0 0 0 0 0.1675 0 0 0 0

0 0 0 0 0 0 0.0611 0 0 0

0 0 0 0 0 0 0 0.0158 0 0

0 0 0 0 0 0 0 0 0.0020 0

0 0 0 0 0 0 0 0 0 0.0001

freq =

1.0e+03 *

7.4446 0 0 0 0 0 0 0 0 0

0 6.6296 0 0 0 0 0 0 0 0

0 0 5.4824 0 0 0 0 0 0 0

0 0 0 4.2398 0 0 0 0 0 0

0 0 0 0 3.0704 0 0 0 0 0

0 0 0 0 0 2.0601 0 0 0 0

0 0 0 0 0 0 1.2437 0 0 0
8 APPENDIX 62

0 0 0 0 0 0 0 0.6322 0 0

0 0 0 0 0 0 0 0 0.2248 0

0 0 0 0 0 0 0 0 0 0.0357

Current plot held

Current plot held

pkdisp1 =

0.0016

pkdisp2 =

0.0055

pkdisp3 =

0.0107

angle_m1 =

0.0053

angle_m2 =

-0.0768
8 APPENDIX 63

angle_m3 =

0.2408

peak_disp_muzzle =

1.0000 2.0000 3.0000

0.0016 0.0055 0.0107

angles =

1.0000 2.0000 3.0000

0.0053 -0.0768 0.2408

v_error_100 =

0.3354 -4.8265 15.1310

>>
8 APPENDIX 64
8 APPENDIX 65

You might also like